Real methods in complex analysis

download Real methods in complex analysis

of 23

Transcript of Real methods in complex analysis

  • 7/23/2019 Real methods in complex analysis

    1/23

    REAL

    PROOFS

    OF

    COMPLEX THEOREMS (AND

    VICE VERSA)

    LAWRENCE

    ZALCMAN

    Introduction.

    It has become

    fashionable

    recently

    to

    argue

    that

    real and

    complex

    variables should be taught together as a unified curriculumin analysis. Now this is

    hardly

    a novel idea,

    as

    a

    quick

    perusal

    of Whittaker

    and Watson's

    Course

    of

    Modern

    Analysis

    or either Littlewood's

    or Titchmarsh's

    Theory

    of Functions (not

    to

    mention

    any

    number

    of cours

    d'analyse

    of the nineteenth

    or twentieth century)

    will

    indicate.

    And, while

    some

    persuasive

    arguments

    can be advanced

    in favor of this

    approach,

    it

    is by

    no

    means

    obvious

    that the advantages

    outweigh the disadvantages

    or,

    for

    that matter,

    that a unified treatment

    offers

    any

    substantial

    benefit to

    the

    student.

    What

    is obvious

    is

    that the two

    subjects

    do

    interact,

    and interact

    substantially,

    often

    in

    a

    surprising

    fashion. These points

    of tangency present

    an instructor

    the

    opportunity

    to

    pose

    (and

    answer)

    natural

    and

    important

    questions

    on basic

    material

    by

    applying

    real

    analysis

    to

    complex

    function

    theory,

    and vice versa. This

    article

    is

    devoted

    to

    several

    such

    applications.

    My

    own

    experience

    in

    teaching suggests

    that the

    subject

    matter discussed

    below

    is

    particularly

    well-suited

    for

    presentation

    in a

    year-long

    first

    graduate

    course

    in

    complex analysis.

    While

    most of this material

    is

    (perhaps

    by definition)

    well

    known

    to the

    experts,

    it is

    not,

    unfortunately,

    a

    part

    of

    the

    common

    culture of

    professional

    mathematicians.

    In

    fact,

    several

    of

    the

    examples

    arose

    in

    response

    to

    questions

    from friends and colleagues. The mathematics involved is too pretty to be the private

    preserve

    of

    specialists.

    Publicizing it is

    the

    purpose

    of the present

    paper.

    1. The Greening

    of

    Morera. One

    of the

    most

    useful

    theorems

    of basic complex

    analysis

    is the

    following result,

    first

    noted

    by

    Giacinto Morera.

    MORERA'S

    THEOREM

    37].

    Let

    f(z)

    be a continuous

    function

    on the domain

    D.

    Suppose

    that

    (1)

    f(z)dz

    =

    0

    for every

    rectifiable

    closed

    curve

    y lying

    in

    D.

    Then f is holomorphic

    in D.

    Morera's

    Theorem enables

    one

    to

    establish the

    analyticity

    of functions

    in

    situations

    where resort to the

    definition

    and

    the attendant calculation of

    difference

    quotients

    would

    lead to

    hopeless complications.

    Applications

    of this sort

    occur,

    for

    instance,

    in the

    proofs

    of the

    Schwarz Reflection Principle and other

    theorems

    on the

    extension

    of

    analytic

    functions.

    Nor

    is its usefulness

    limited

    to this

    circle of ideas;

    the

    important

    fact

    that

    the uniform limit of

    analytic

    functions

    is

    again

    analytic

    is an

    immediate consequence (observed already by Morera himself, as well as by Osgood

    [39],

    who had

    rediscovered

    Morera's

    theorem).

    115

  • 7/23/2019 Real methods in complex analysis

    2/23

    116

    LAWRENCE

    ZALCMAN

    [February

    Perhaps

    surprisingly,

    the

    proofs

    of Morera's theorem found

    in complex analysis

    texts

    all follow a single pattern.

    The hypothesis

    on

    f

    insures

    the existence of a

    single-valued

    primitive

    F of

    f,

    defined

    by

    rz

    (2)

    F(z)

    f(C)dC.

    0

    Here

    zo

    is some

    fixed

    point

    in

    D

    and

    the

    integral

    is taken over

    any

    rectifiable

    curve

    joining

    zo

    to z. The function

    F is

    easily

    seen to be

    holomorphic

    in

    D,

    with

    F'(z)

    =

    f(z);

    since the

    derivative

    of a

    holomorphic

    function is

    again

    holomorphic,

    we

    are done.

    Several remarks

    are in order concerning

    the proof

    sketched above.

    First

    of all,

    the assumption

    that

    (1)

    holds

    for all rectifiable closed

    curves

    in D is much too

    strong.

    It

    is

    enough,

    for

    instance,

    to assume

    that

    (1)

    holds for all

    closed curves

    consisting

    of a finite number of straight line segments parallel to the coordinate axes; the

    integration

    in

    (2)

    is then effected

    over a

    (nonclosed)

    curve

    composed

    of such

    segments,

    and the proof proceeds

    much

    as before.

    Second,

    since

    analyticity

    is a

    local

    property,

    condition

    (1)

    need

    hold

    only

    for an

    arbitrary neighborhood

    of each

    point

    of

    D;

    that

    is,

    (1)

    need

    hold

    only

    for small curves.

    Finally,

    the

    proof requires

    the

    fact

    that

    the derivative

    of

    an

    analytic

    function is

    again

    analytic.

    While this

    is a

    trivial con-

    sequence

    of the

    Cauchy integral

    formula,

    it can be

    argued

    that

    that

    is an

    inappropriate

    tool for

    the problem

    at

    hand;

    on the other

    hand,

    a

    proof

    of this fact

    without

    complex

    integration

    is

    genuinely

    difficult

    and

    was,

    in

    fact,

    only

    discovered

    (after many years

    of

    effort)

    in

    1961

    [44],

    [10], [46].

    There is an additional

    defect

    to the

    proof,

    and that is that

    it

    does not

    generalize.

    Thus, it was more

    than

    thirty years

    after

    Morera

    discovered his

    theorem that

    Torsten

    Carleman

    realized

    the

    result

    remains

    valid

    if

    (1)

    is

    assumed

    to

    hold

    only

    for

    all

    (small) circles

    in D.

    It is

    an

    extremely

    instructive

    exercise

    to

    try

    to

    prove

    Carleman's

    version

    of Morera's

    theorem

    by mimicking

    the

    proof given

    above. The

    argument

    fails

    because it cannot

    even

    be started:

    the

    very

    existence of

    a

    single-valued

    primitive

    is in

    doubt.

    This

    leads

    one to

    try

    a different

    (and

    more

    fruitful) approach,

    which

    avoids the use of primitives altogether.

    Suppose

    for

    the

    moment

    that

    f

    is a smooth

    function, say

    continuously

    dif-

    ferentiable.

    Fix

    zo

    e

    D and

    suppose

    (1)

    holds

    for

    the circle

    Fr(ZO)

    of radius

    r,

    centered

    at

    zo.

    Then,

    by

    the

    complex

    form of

    Green's.

    theorem

    0 f(z)dz

    =

    2i

    jj

    dxdy,

    Z O )

    Z O o f

    where

    4(Zo)

    is

    the

    disc

    bounded

    by

    FU(zO)

    and

    Of/lz

    =

    2

    (3f/Ox

    +

    i

    Of/Dy).

    (There's

    no cause

    for

    panic

    if the

    0/Di operator

    makes

    you

    uneasy

    or

    you

    are not

    familiar

    with the

    complex

    form of Green's

    theorem; just

    write

    f(z)

    =

    u(z) + iv(z),

    dz

    =

    dx

    + idy,

    and

    apply

    the

    usual version

    of Green's theorem

    to the real and

    imaginary

    parts

    of

    the

    integral

    on

    the

    left.) Dividing by

    an

    appropriate

    factor,

    we

  • 7/23/2019 Real methods in complex analysis

    3/23

    1974]

    REAL

    PROOFS OF COMPLEX THEOREMS

    117

    have

    7tr2 t

    (:) jedxdy =

    0;

    i.e., the average of the continuous function Of/0 over the disc

    AF(zo)

    equals 0.

    Make

    r

    -+

    0 to obtain (Of

    O)

    (z0)

    =

    0.

    Since

    this holds

    at

    each

    point

    zo

    e

    D,

    Of

    Oz=

    0

    identically

    in

    D. Writing this

    in real

    coordinates,

    we see

    that

    ux

    =

    vp,

    uY

    =

    -vx

    in D; thus the Cauchy-Riemann equations

    are

    satisfied

    and

    f

    is

    analytic.

    Notice that we did not need to assume

    that

    (1)

    holds

    for

    all

    circles in

    D or

    even

    all small

    circles;

    to

    pass

    to

    the

    limit it was

    enough

    to

    have,

    for each

    point

    of

    D, a

    sequence of circles shrinking to that point. Moreover, since f

    has

    been assumed

    to

    be continuously differentiable, it is sufficient to prove that

    af/iz

    vanishes on

    a

    dense set. Finally, and most important, the fact

    that our curves were circles

    was

    not used

    at all Squares, rectangles, pentagons,

    ovals

    could have been used

    just

    as well. To conclude that

    (Of/li)(zo)

    =

    0, all

    we

    require

    is that

    (1)

    should

    hold

    for a

    sequence

    of

    simple closed

    curves

    y

    that accumulate

    to

    zo

    (zo

    need

    not

    even

    lie

    inside or on the y's) and that the curves involved

    allow

    application

    of

    Green's

    theorem. It is enough, for instance, to assume that the curves are piecewise con-

    tinuously differentiable.

    To summarize, we have shown that Green's theorem yields in a simple fashion

    a

    very general and particularly appealing version

    of

    Morera's

    theorem for C'

    func-

    tions. It may reasonably be asked at this point if the proof of Morera's theorem

    given above can be modified to work for functions

    which

    are

    assumed

    only

    to

    be

    continuous. That is the subject of the next section.

    2.

    Smoothing. Let +(z) be a real valued function defined

    on

    the

    entire

    complex

    plane which satisfies

    (a)

    0(z)

    >

    09

    (b)

    ff

    q(z)dxdy

    =

    1,

    (c)

    4

    is

    continuously differentiable,

    (d) 0(z) =

    0

    for Iz I _ 1.

    It is trivial to construct such functions; we can even require

    /

    to be infinitely dif-

    ferentiable and to depend only on

    I

    z

    ,

    but these properties will not be required in the

    sequel. Set, for

    e

    > 0,

    4,(Z)

    =

    C-20(z/E).

    Then, clearly, 4? satisfies (a) through (c)

    above

    and

    /E

    vanishes off

    j

    I

    O

    >

    n

    2

    so that f

    is

    univalent.

    The

    second

    ingredient we

    need is the celebrated Hadamard

    gap

    theorem.

    HADAMARD

    GAP THEOREM.

    Let f(z)

    =

    U=O

    a

    znk

    have A

    as its

    disc

    of con-

    vergence. If

    nk+link

    _

    qfor

    some q > 1 and all

    large k,

    thenf

    has

    F

    as

    its

    natural

    boundary; that is, f cannot be continued analytically across any subarc of F.

    The

    beautiful proof

    of this theorem due to L. J. Mordell

    ([36],

    cf.

    [54,

    p. 223])

    should be

    standard fare in

    graduate

    courses in complex

    analysis.

    The

    construction of

    the

    required function is

    now almost trivial.

    We

    choose the

    sequences

    {ak}

    and {nk}

    to satisfy

    (a)

    ao

    =

    no

    =

    1,

    (b)

    zko1

    nkf

    ak| 0 [61]. Even this condition is not necessary;

    in

    fact,

    Denjoy [11]

    has constructed an arc which

    is

    the graph of a function and

    which

    has the

    required property.

    6.

    Blowing up

    the

    boundary. Questions involving length and area arise in con-

    formal

    mapping

    as well.

    A

    conformal

    map, being

    analytic,

    must map sets of

    zero

    area to

    sets

    of zero area; however, distortion at the boundary is an a priori possibility.

    Writing

    A

    u

    F

    =

    {z:

    z

    f

    ?

    1}

    as

    before,

    let us assume

    that

    the

    univalent function

    f(z) maps

    A

    conformally

    onto

    the Jordan region D. According to the Osgood-

    Taylor-Caratheodory theorem,

    f

    extends to a homeomorphism of A u F onto

    D

    u OD.

    (Proofs

    of

    this

    important result, announced by Osgood [65] and proved

    independently by Osgood

    and

    Taylor [66, p. 294], and

    Caratheodory

    [6], [7] are

    available

    in

    [9, pp. 46-49] and [24, p. 129-134]. The reader will find a comparison

    of the

    treatments

    in

    these

    references particularly instructive in the matters of style

    of

    exposition

    and

    attention

    to

    detail.) In case OD is rectifiable, a theorem of the

    Riesz brothers

    [47]

    insures that

    f

    and

    f

    -'

    preserve sets of zero length (

    =

    Haus-

    dorff

    one-dimensional

    measure). When

    OD

    fails to be

    rectifiable,

    however, all hell

    breaks loose.

    In

    particular,

    a

    subset of

    OD

    having positive area may correspond to

    a subset of F having zero Lebesgue (linear) measure For the construction, we need

    an

    important result from plane topology.

    MOORE-KLINE

    EMBEDDING THEOREM

    [351.A necessary and sufficient condition

  • 7/23/2019 Real methods in complex analysis

    10/23

    124

    LAWRENCE ALCMAN

    [February

    that

    a

    compact

    set K

    (

    C

    should

    lie

    on

    a

    simple

    Jordan arc is that

    each

    closed

    connected subset

    of

    K

    should be either a

    point

    or

    a

    simple

    Jordan arc with

    the

    property that

    K

    -

    y

    does

    not

    accumulate at

    any point of

    7,

    except

    (perhaps) the

    endpoints.

    Now

    let K be

    a

    Cantor set

    having positive

    area;

    K

    may

    be

    realized,

    for

    instance

    as the product of

    two linear Cantor

    sets,

    each

    of which has

    positive

    linear

    measure.

    Construct

    countably

    many disjoint

    simple

    Jordan arcs

    Jn

    c

    C

    \ K

    such that

    the

    sequence

    {Jn}

    accumulates

    at

    each

    point

    of K and at no other

    points

    of

    C and

    with

    the

    additional

    property that

    if

    zo

    e

    C\

    (K

    tu

    {Jn})

    =

    R and

    z e

    K,

    then

    any

    arc

    from

    zo

    to

    z

    which lies,

    except

    for its final

    endpoint,

    in

    R

    must

    have

    infinite

    length.

    By

    the Moore-Kline

    embedding

    theorem,

    we

    may pass

    a

    simple

    closed Jordan

    curve J through K

    U

    {JnJ.

    Let

    f

    be a conformal

    map

    from

    A to

    D,

    the

    domain

    bounded

    by

    J. Then

    f extends to a

    homeomorphism

    from

    F

    to J.

    Let

    S

    =

    f

    -'(K).

    That

    S has zero linear

    measure

    follows at once from the

    following

    theorem,

    due

    to

    Lavrentiev.

    THEOREM.Let

    f be a

    conformal

    homeomorphism

    of

    Au

    Fu

    onto

    the Jordan

    domain D

    UJ.

    If

    S

    c

    J is

    not

    rectifiably

    accessible

    from

    D

    then

    f

    '(S)

    c

    F

    has

    zero

    measure.

    Proof.

    Since

    D is a

    bounded domain, its

    area,

    given by the

    expression

    f fA

    f '(z)

    I2dxdy,s

    finite. Thus

    2n

    1

    f

    f'(reio)

    I

    rdrdO

    2ir 1

    1

    2

    2r2

    1

    112

    ?

    (Jj

    k O

    rdrdO

    If'(reE?) r

    drdO

    0. The

    first

    person to

    show that

    a set of

    capacity zero

    on

    F

    could correspond

    under a

    conformal

    mapping to a set having positive area was Kikuji Matsumoto [33]. He actually

    proved

    (what is implicit in

    the

    above discussion)

    that for each

    totally

    disconnected

    compact subset K

    of the plane

    there

    exists a Jordan

    domain D with

    boundary

    J

    D

    K

    such

    that K

    corresponds

    under

    conformal

    mapping to a set of

    capacity

    zero

  • 7/23/2019 Real methods in complex analysis

    11/23

    1974]

    REAL PROOFS OF COMPLEX THEOREMS

    125

    on

    F. The discussion here

    (in

    particular,

    the ingenious proof of the central result)

    is based

    on an idea of Walter Schneider

    [51].

    The compression

    of

    the

    boundary

    of

    the unit disc

    presents

    greater difficulties.

    Lavrentiev, however, has shown that a set of positive measure on F may be mapped

    onto a set of zero length under a conformal mapping of Jordan

    domains

    [30].

    A

    more recent construction is due to McMillan and Piranian

    [32].

    7. Absolute convergenceand uniformconvergence.

    Conformal

    mapping

    techniques

    are also useful in constructing examples concerning the convergence of

    power

    series and Fourier series. Below we offer some simple but instructive

    examples.

    The first example

    of a

    power

    series which

    converges

    uniformly

    but not

    absolutely

    on

    the closed unit disc was

    given by Fejer [15],

    cf.

    [25,

    vol.

    1,

    p. 122].

    The

    following

    geometric example, due to Gaier [68] and rediscovered by Piranian (see [1, pp. 289,

    314]),

    is

    particularly appealing.

    Let D be the region of figure 1, a triangle from

    FIG.

    1

    which

    wedges have been removed

    in such a way that the vertex at z

    =

    1

    is not

    rectifiably accessible from the

    interior of D. Since D is a Jordan

    region, any conformal

    map

    of

    A

    =

    {z:

    z

    (x)

    e

    Y, q$(xk)

    e

    J

    (k

    =

    1, 2,

    )

    and

    5

    is

    closed

    under

    linear

    combina-

    tions. Since

    (by

    hypothesis) x

    e

    5, each

    polynomial vanishing

    at the origin

    belongs

    to

    S.

    The proof

    depends on a

    simple lemma

    concerning the

    approximation

    of

    functions.

    LEMMA.

    Let

    0(x)

    satisfy

    (a). Suppose that

    for

    each

    E

    >

    0

    there

    exist

    poly-

    nomials

    pl(x),

    p2(X) such that

    pi(O)

    =

    0,

    pi(l)

    =

    1 (i

    =

    1,2) and

    P(x)

    N. (The

    inner circle

    contracts,

    the

    outer circle

    expands, and

    the notch

    gets

    thinner and rotates,

    tending

    toward

    but

    never

    reaching its

    limiting line.)

    The

    function

    0? z cKn

    gn(Z)

    { K

    clearly

    extends to

    a function

    analytic

    in a

    neighborhood

    of the (disconnected)

    set

    Kn

    u

    {0}.

    Since

    this set does

    not separate the

    plane,

    g may be

    approximated

    uni-

    formly (to

    within

    1/n, say) on

    Kn

    u

    {0}

    by

    a polynomial

    pn.

    These polynomials

    clearly

    satisfy

    (11).

    12. Miscellany.

    The interactions

    between real

    and complex

    analysis

    are by no

    means limited to the areas mentioned above. To keep the discussion within manage-

    able

    limits,

    we have restricted

    ourselves

    to

    (a

    subset of)

    those

    applications,

    examples,

    and

    aspects

    of the theory

    that have

    not

    found sustained

    treatment

    in the "popular"

    literature of texts

    and survey

    articles.

    Subjects which

    are

    treated elsewhere

    at ade-

    quate

    length but

    which deserve

    mention

    here by

    virtue of their

    interdisciplinary

    nature include

    the following:

    (a) The evaluation

    of real

    integrals

    and sums

    by

    residue techniques.

    This is

    surely

    one of the

    most striking

    applications

    of complex

    function

    theory to real

    analysis. Fortunately,

    any

    good text on

    complex

    analysis will

    contain a

    fairly detailed

    discussion.

    (b)

    Complex methods

    in harmonic

    analysis.

    This is a substantial area,

    which

    includes topics

    as

    diverseas interpolation

    theorems

    (see,

    for instance,

    [28, pp. 93-98])

    and

    theorems

    of Paley-Wiener

    type [43].

    Two

    of the most

    attractive recent

    texts

    in

    harmonic analysis

    [13], [28]

    devote

    whole chapters

    to this

    aspect

    of the theory.

    Further

    developments are

    discussed

    in

    the survey article

    of

    Weiss [57].

    (c)

    Functional

    analysis. Complex

    variable

    methods

    appear

    here

    perhaps

    most

    notably in the construction

    of

    functional

    calculi

    for operators

    on Hilbert

    space

    or

    Banach space. The applications to commutative Banach algebras are particularly

    substantial;

    indeed, parts

    of this

    last-named

    subject

    are

    virtually

    coextensive

    with

    certain

    aspects

    of

    several

    complex

    variable

    theory.

    For

    further

    references,

    see

    [17]

    and

    [58].

    In the

    opposite

    direction,

    techniques

    of

    functional

    analysis

    can be

    used

  • 7/23/2019 Real methods in complex analysis

    20/23

    134 LAWRENCE ZALCMAN

    [February

    to

    establish many results in

    function theory; this is the programme of [48].

    Finally,

    an

    honest partnership between

    complex variables and functional analysis occurs

    in

    the

    study of certain Banach spaces of analytic functions, especially HP

    spaces

    [12],

    [26].

    (d) Function theoretic methods in

    differential

    equations. Complex

    methods

    occur rather naturally in the

    study of ordinary differential equations [8].

    Their

    appearance in the study of

    partial differential equations is perhaps more

    surprising.

    Yet there are substantial applications, and more than one book [2], [19]

    has been

    devoted to this area. Further applications of function theory to

    problems in partial

    differential equations will be

    found in [18]. In a rather different direction, the theory

    of

    linear partial differential equations with constant coefficients is

    intimately con-

    nected with the study of certain spaces of entire functions of several

    complex varia-

    bles; see [14] for an exhaustive treatment.

    13. Monodromy. No excursion onto the bypaths of complex analysis would be

    complete without some

    mention of the monodromy theorem.

    MONODROMY THEOREM.

    Let D be a simply connected domain and let

    f(z)

    be

    analytic in a neighborhood of

    zo

    e

    D. Then if f(z) can be continued

    analytically

    from

    zo

    along every path lying in D, the continuation

    gives]

    rise to a

    single-valued

    unction analytic on all

    oJ

    D.

    A more general version states that analytic continuation along paths is a homo-

    topy invariant; see, for

    instance, [53]. Like the reflection principle, the monodromy

    theorem is an essential

    ingredient in the short proof of Picard's little theorem;

    in

    its

    extended

    form, it is the

    central result in the subject of analytic continuation.

    Yet

    no

    theorem of basic complex analysis is more abused or less understood.

    Indeed,

    it

    has been

    misapplied more than once even by mathematicians of

    the

    first

    rank

    (and

    specialists in complex analysis,

    at that ). One may speculate that

    a source

    of at

    least

    some of the confusion surrounding this result is the essentially

    topological,

    rather than function-theoretic,

    nature of the theorem.

    The sort of error into which one may lapse is best indicated by an explicit example.

    Let

    D

    be a simply connected

    domain and f a function analytic

    in D which satisfies

    f '(z)

    0

    0 on D.

    Suppose R

    =

    f(D)

    is

    also

    simply

    connected.

    Question:

    Must

    f

    be

    univalent

    (one-one)? An affirmativeanswer may be found in [56, p. 243] and

    in

    other

    references as well. The

    argument is as follows. At each point w0

    E

    R one may

    define

    a

    local inverse fJt1(w) of f,

    analytic in a neighborhood of w0. Since R

    is

    simply

    connected, the totality of these functions defines a single-valued analytic

    function

    f

    -1

    on

    R, which is a global

    inverse for f. Thus f must be univalent.

    Note

    further

    that

    the

    simple-connectivity of

    D

    is quite extraneous to the demonstration.

    Unfortunately,

    the

    argument given above

    is

    altogether incorrect,

    since

    the

    essential hypothesis of the

    monodromy theorem, that analytic continuation

    be

    possible along every path

    in

    R, has not been verified.

    Can

    the proof

    be

    salvaged?

  • 7/23/2019 Real methods in complex analysis

    21/23

    1974]

    REAL

    PROOFS

    OF

    COMPLEX

    THEOREMS

    135

    The answer

    is no. In fact,

    consider the

    function f(z)

    =

    f

    e2

    dC.

    This

    f

    is analytic

    (entire) on

    the simply

    connected domain C

    and f'(z) _

    eZ2

    is

    nowhere

    zero. Clearly,

    f is

    not univalent. So

    it suffices to prove

    that

    f(C)

    is simply

    connected. We claim

    f(C)

    =

    C. Indeed, suppose f(z) # w. Since

    eZ2

    is an even function, f(z) is odd,

    so

    that

    if

    f

    fails to take on the

    value

    w

    it

    also misses the value

    -

    w.

    If w

    =A

    ,

    this

    contradicts

    Picard's

    (small) theorem.

    Since

    f(O)

    =

    0,

    f takes on

    every

    value in

    the

    complex plane.

    For

    D

    =

    C, any

    function

    which satisfies f'(z)

    0

    0 must be

    transcendental

    and hence must (by

    Picard's

    theorem) take on most values

    infinitely

    often.

    One

    can,

    however, construct a

    non-univalent, locally univalent function

    mapping

    the

    disc

    A

    =

    {z:

    I

    z

    I